摘要
Open AccessCCS ChemistryRESEARCH ARTICLE14 Jul 2022Chiral CuxCoyS Supraparticles Ameliorate Parkinson’s Disease Baimei Shi†, Aihua Qu†, Weiwei Wang, Meiru Lu, Zhuojia Xu, Chen Chen, Changlong Hao, Maozhong Sun, Liguang Xu, Chuanlai Xu and Hua Kuang Baimei Shi† International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 †B. Shi and A. Qu contributed equally to this paper.Google Scholar More articles by this author , Aihua Qu† International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 †B. Shi and A. Qu contributed equally to this paper.Google Scholar More articles by this author , Weiwei Wang International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author , Meiru Lu International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author , Zhuojia Xu International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author , Chen Chen International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author , Changlong Hao International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author , Maozhong Sun International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author , Liguang Xu *Corresponding author: E-mail Address: [email protected] International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author , Chuanlai Xu International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author and Hua Kuang International Joint Research Laboratory for Biointerface and Biodetection, Jiangnan University, Wuxi, Jiangsu 214122 State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi, Jiangsu 214122 Google Scholar More articles by this author https://doi.org/10.31635/ccschem.021.202101107 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesDownload CitationsTrack Citations ShareFacebookTwitterLinked InEmail Protein aggregation causes alpha-synuclein (α-syn) to change from its original physiological role to a pathological state, which is a potential pathogenic mechanism in Parkinson’s disease (PD). Chiral l/d-CuxCoyS supraparticles (l/d-SPs) with a circular dichroism value of 35 mdeg at 805 nm were fabricated using a simple wet-chemical method. The l/d-SPs prevented the α-syn monomers from forming fibrils and triggered the α-syn fibrils to turn into monomers under 808 nm near-infrared (NIR) light. In living MN9D cells, d-SPs reduced cellular damage, neuronal functional deficits, and neuron loss caused by α-syn fibrils after NIR spectroscopy treatment within 10 min to prevent α-syn aggregation. Significantly, the reactive oxygen species produced by d-SPs were 1.42 times higher than those produced by l-SPs. In vivo experiments showed that d-SPs had a protective effect on neuron damage caused by α-syn aggregate deposition, reduced the symptoms in a mouse model of PD, and restored cognitive ability. After NIR light treatment, the amount of α-syn in a mouse model of PD decreased by more than 67.5%. At the same time, d-SPs gradually decomposed into small nanoparticles within 60 days and were excreted through the blood–brain barrier. This discovery paves the way for the treatment of neurodegenerative diseases using chiral SPs under NIR light irradiation. Download figure Download PowerPoint Introduction Parkinson’s disease (PD) is a common neurodegenerative disease and mainly manifests as static tremor, bradykinesia, myotonia, and postural gait disorders.1 The abnormal filamentous aggregation of alpha-synuclein (α-syn) may be related to the pathological dysfunction of PD.2 Extracellular α-syn can also enter recipient cells, and aggregates are more internalized and toxic than monomers. Therefore, clearance of α-syn aggregates in extracellular fluid is very important for treatment of PD. Chirality exists widely in nature and plays an important role in biological systems.3 Due to the huge application foreground in biosensing and biomedicine, more and more chiral inorganic nanomaterials have been prepared.4–6 It has recently been discovered that by manipulating the chirality of nanomaterials, the interaction between the material and the cell can be affected.7–11 Inorganic supraparticles (SPs)12–14 formed with inorganic nanoparticles have advantages compared to solid nanoparticles of the same size because of their high surface area and porosity. Due to their structural specificity and functional superiority, SPs have great application potential in the fields of biocatalysis, bioimaging, and biosensing.15–21 Chiral SPs with various morphologies can be synthesized by adjusting ligand types, elemental composition, ionic strength, pH values, and other conditions. Some studies have shown that chiral SPs with porous structure can improve their catalytic activity20 and can be used for drug delivery.16 At present, nanomaterials including metal nanoparticles22 and graphene quantum dots23 are used to inhibit or eliminate α-syn in the treatment of PD diseases, and there is no research on the treatment of PD using chiral SPs. Here, we aimed to synthesize chiral CuxCoyS SPs (l/d-SPs) with l- or d-penicillamine (l/d-Pen) as ligands, and found that d-SPs possessed higher disaggregation ability of α-syn fibrils than l-SPs under near-infrared (NIR) illumination for 10 min. To explore the potential mechanism, reactive oxygen species (ROS) produced by l/d-SPs under NIR illumination were characterized. The d-SPs under NIR illumination generated a higher number of ROS than l-SPs and then effectively dissociated the α-syn aggregates, reducing the neuronal toxicity caused by these aggregates. It is worth noting that both in vitro (MN9D cells) and in vivo (PD model mice) experiments successfully verified the therapeutic effect by d-SPs. Experimental Methods Under the heating condition protected by nitrogen, l/d-Pen reacted with Cu(ClO4)2·H2O, CoCl2·6H2O, and thioacetamide to prepare chiral CuxCoyS SPs. Transmission electron microscopy (TEM), high-resolution TEM (HRTEM), dynamic light scattering (DLS), energy-dispersive X-ray (EDX), and circular dichroism (CD) were used to characterize the material properties. TEM, CD, and electrophoresis were used to explore the degradation of α-syn fibrils under 808 nm light irradiation. Then, MN9D cells were treated with fluorescein isothiocyanate (FITC)-conjugated α-syn, and the disaggregation of syn protein at the cell level under the action of NIR was observed by immunofluorescence staining. Subsequently, the causes of l/d-SPs to disaggregate α-syn were explored, X-ray photoelectron spectroscopy (XPS) was used to analyze l/d-SPs valence changes, and isothermal titration calorimetry (ITC) was applied for affinity determination. Finally, d-SPs were injected into the brains of PD model mice, and this was followed by the light treatment. We determined the treatment of the PD mice by observing changes in their behavior in the water maze and immunohistochemical and immunofluorescence sections of their brains. All animal treatment and maintenance protocols were approved by the Institutional Animal Care and Use Committee of Jiangxi University. More experimental details are available in the Supporting Information. Results and Discussion Synthesis and characterization of chiral l/d-CuxCoyS SPs To prepare chiral CuxCoyS SPs, l/d-Pen was chosen as the chiral source, copper(II) perchlorate hexahydrate as the copper source, cobalt(II) chloride hexahydrate as the cobalt source, and thioacetamide as the sulfur source (Scheme 1). The average diameters of d-SPs and l-SPs measured by TEM were 87 ± 6 nm and 89 ± 4.5 nm, which were consistent with the hydrodynamic size of SPs measured by DLS (as ∼88 and 90 nm for d- and l-SPs, respectively) (Figure 1a and Supporting Information Figures S1a and S2). A representative HRTEM image showed that l/d-SPs were porous structures with an average pore diameter of 0.19 nm formed by self-assembly of small nanoparticles (Figure 1b and Supporting Information Figure S1b), which was also confirmed by high-angle annular dark-field scanning TEM (HAADF-STEM) images (Figure 1c and Supporting Information Figure S1c). Scheme 1 | Schematic of (a) l/d-Pen-modified l/d-CuxCoyS SPs. (b) Illustration of the inhibition and disassembly effects of d-SPs on α-syn aggregation and mitigation of potential neurotoxicity in a PD mice model. Download figure Download PowerPoint The well-defined lattice spacing of 0.190 and 0.191 nm corresponded to the (110) plane of CuS and the (422) plane of Co3S4, respectively (Figure 1b). The elemental mapping of copper, cobalt, sulfur, nitrogen, and oxygen was analyzed, which was consistent with the results of EDX spectra (Figure 1d and Supporting Information Figures S1d and S3). The composition of the SPs was characterized by X-ray diffraction (XRD). The main peaks of the SPs matched well with the standard pattern of CuS (JCPDS no. 24-0060) and Co3S4 (JCPDS no. 02-0825) ( Supporting Information Figure S4). Figure 1 | (a) TEM image and SAED pattern, (b) HRTEM image, (c) HAADF-STEM image, and (d) energy-dispersive X-ray spectroscopy (EDS) mapping images of d-SPs. High-resolution (e) Cu 2p and (f) Co 2p XPS spectrum of d-SPs. (g) CD spectra of l/d-SPs and l/d-SPs at 190–1000 nm. (h) The absorbance spectra of l/d-SPs and l/d-SPs at 190–1000 nm. Download figure Download PowerPoint XPS of a Cu 2p scan for SPs revealed two main peaks at 955.6 and 932.3 eV; these corresponded to Cu 2p1/2 and Cu 2p3/2, respectively. The Cu 2p3/2 peak could be fitted to two peaks with binding energies of 932.3 and 934.5 eV. These demonstrated that the valence of Cu was a mixture of the 1+ and 2+ states (Figure 1e and Supporting Information Figure S1e). According to the fitting peaks of the Co 2p3/2 peak and the Co 2p1/2 peak, the valence of Co was a mixture of the 2+ and 3+ states (Figure 1f and Supporting Information Figure S1f). These results suggested that the copper and cobalt in SPs exhibited multiple covalent states. In addition, zeta potential was −16.23 ± 0.65 mV for l-SPs and −18.7 ± 1.56 mV for d-SPs ( Supporting Information Table S1). The optical properties of the SPs were measured by UV–vis absorption and CD spectroscopy in the range of 190–1000 nm. l- and d-SPs displayed mirror-symmetrical CD responses in the range of 190–1000 nm and a CD value of ±35 mdeg at 830 nm (Figure 1g), corresponding to the absorption spectra (Figure 1h). Interaction of α-syn fibrils/monomers with l/d-SPs under NIR illumination Multiple parameters such as the irradiation time, irradiation power, and concentration of d-SPs were optimized for subsequent experiments ( Supporting Information Figures S5–S7). TEM was used to characterize the changes in α-syn polymerization states (10 μM) with l/d-SPs (100 μg/mL) in vitro with different NIR light (200 mW/cm2) irradiation time. As the time increased from 0 to 10 min, the α-syn fibrils gradually dissociated into short fragments, and finally changed to oligomers and monomers (Figure 2a). As further confirmed by native polyacrylamide gel electrophoresis (PAGE), d-SPs gradually broke the fibrils into monomers under 808 nm NIR light illumination for 10 min. In contrast to d-SPs, partial α-syn fibers could not be dissociated by l-SPs under the same irradiation conditions (Figure 2b). The CD spectra were used to characterize the change of fibrils after interaction with the d-SPs under NIR light irradiation. As shown in Figure 2c, the CD peak at 220 nm which corresponds to the secondary structure (β-sheet) decreased dramatically within 10 min, indicating a decrease in the component in the fibrils.23 On the contrary, when l-SPs were incubated with the α-syn fibers for the same illumination time, the fibrils retained the β-sheet structure (Figure 2c). These results indicated that, under the same conditions, the ability to dissociate fibrils by d-SPs was better than that of l-SPs. Meanwhile, the CD spectra of the reaction system at 400–1000 nm did not change significantly, indicating the good stability of chiral SPs ( Supporting Information Figure S8).7 Figure 2 | (a) TEM images of preformed α-syn fibrils treated with l/d-SPs, and under NIR irradiation (200 mW/cm2) for different times. (b) Native-PAGE analysis of α-syn with different treatments [(1) marker, (2) α-syn monomers without treatment, (3) α-syn fibrils + d-SPs + NIR 0 min, (4) α-syn fibrils + d-SPs + NIR 3 min, (5) α-syn fibrils + d-SPs + NIR 5 min, (6) α-syn fibrils + d-SPs + NIR 10 min, (7) α-syn fibrils + l-SPs + NIR 10 min]. (c) CD spectra of α-syn fibrils treated with l/d-SPs, and under NIR irradiation (200 mW/cm2) at 190–250 nm. (d) TEM images and (e) the corresponding CD spectra of α-syn monomers or monomers with different treatments (monomer only, l/d-SPs, l/d-SPs under NIR irradiation). All α-syn monomer samples (100 μM) seeded with α-syn fibrils (10 μM) before incubation at 37 °C with shaking at 600 rpm for 24 h, except the monomer without treatment. Mono in this figure means monomers. Download figure Download PowerPoint The role of SPs in the inhibition of α-syn fibrils under the same NIR light conditions was also studied. As shown in Figure 2d, the untreated α-syn monomer sample remained in the α-syn monomer state after incubation at 37 °C for 24 h. The α-syn monomer samples seeded with α-syn fibrils (10 μM) transformed into long fibrils after incubation under the same conditions. The control experiments showed that NIR irradiation for only 10 min cannot prevent the formation of α-syn fibrils ( Supporting Information Figure S9). And only d- or l-SPs without NIR irradiation did not completely prevent the formation of α-syn fibrils. Noticeably, there was no fibril formation when the sample was treated with both d-SPs and NIR irradiation, while a small number of fibrils formed using l-SPs as the inhibition agents under the same illumination conditions (Figure 2d). CD spectra were also used to confirm the inhibitory effect of the chiral SPs. As shown in Figure 2e, the monomers turned into fibrils in the absence of SPs. Meanwhile the treatment with l-SPs and NIR light totally inhibited the formation of α-syn aggregates. In the range of 400–1000 nm, the peak position and intensity of the CD spectra did not change, indicating that the l/d-SPs retained their original properties ( Supporting Information Figure S10). l/d-SPs prevent MN9D neuronal cells pathology induced by α-syn aggregation To evaluate the potential bioapplication of l/d-SPs, we conducted cell viability tests to study their biocompatibility. Six different concentrations of l/d-SPs were incubated with MN9D cells for 12 h, and cell viability was measured with the cell counting kit-8 (CCK-8) assay. When the concentration of l/d-SPs was increased, cell viability was unchanged up to the concentration of 100 μg/mL of l/d-SPs. Thus, 100 μg/mL of l/d-SPs was chosen for subsequent experiments. At this concentration, different incubation times had no obvious influence on cell viability ( Supporting Information Figure S11). To visualize the effect of α-syn fibrils on MN9D cells, α-syn fibrils were labeled with FITC ( Supporting Information Figure S12). The time of l/d-SPs entry into the MN9D cells (12 h) was optimized to maximize intracellular concentration ( Supporting Information Figures S13 and S14). At the same time, the intensity (200 mW/cm2) and time (10 min) of NIR irradiation were optimized ( Supporting Information Figures S15–S22). As shown in Figure 3a, after treating the cells with α-syn fibrils, a large number of fibers adhered to the cell membrane surface and then caused changes in cell growth. In the α-syn + NIR group, the adherent α-syn fibrils on the cell surface did not change significantly, indicating that NIR light illumination alone cannot dissociate the fibrils. Just adding d-SPs partially reduced fibrils on the cell surface. When both d-SPs and NIR light were simultaneously present, most of the α-syn fibrils adhering to the surface disappeared after 10 min of NIR light. The results of the l-SPs on the dissociation of fibrils are shown in Supporting Information Figure S23. As expected, MN9D cells in the group containing d-SPs showed better ability to dissociate the α-syn fibrils attached to cell membranes under NIR light illumination. To monitor the effects of different treatments in more detail, we observed changes in neurite outgrowth length (Figure 3b and Supporting Information Figure S24a). MN9D cells treated with l/d-SPs showed a mean length of neurite growth of up to ∼80 μm during culture under 808 nm NIR light. In contrast, most of the cells in the other treatment groups showed short neurite processes and rounded cell morphology. The expression level of Map2 (the mature neuron marker24) was also higher in the cotreatment groups with l/d-SPs and 808 nm NIR light (Figure 3c and Supporting Information Figure S24b). Concurrently, we evaluated the possibility of l/d-SPs preventing α-syn monomers forming fibrils. As shown in Supporting Information Figures S25 and S26, as expected, under the conditions of 200 mW/cm2 and 10 min illumination of d-SPs, no α-syn fibrils were produced, and similar results were obtained for l-SPs. Figure 3 | (a) Confocal images of MN9D cell incubated with α-syn fibrils or under different treatments. The concentration of α-syn fibrils, labeled with FITC, was 5 μM (α-syn-FITC/α-syn = 1/10). Red: Map2 (as mature neuronal biomarker), green: α-syn fibrils labeled with FITC (α-syn-FITC), white: d-SPs, blue: 4′,6-diamidino-2-phenylindole (DAPI) for the nuclei. Scale bar, 20 μm. (b) Neurites mean length of MN9D with different treatments. (c) Fluorescence intensity of Map2 for MN9D under different treatments (NIR, d-SPs, and d-SPs under NIR irradiation). Data are presented as mean ± s.d. (n = 5). Download figure Download PowerPoint Assessment of the mechanism of l/d-SPs facilitating the disaggregation of α-syn fibrils Several lines of evidence show that ROS may play a pivotal role in the process of protein oxidation and disaggregation. We used a 2′,7′-dichlorodihydrofluorescein (DCFH) probe to monitor the generation of ROS by the mixture of d-SPs and α-syn fibrils during NIR irradiation. It was found that the content of ROS in the mixture increased with the extension of illumination time. During the illumination period, ROS production after 10 min incubation was highest, and ROS production in d-SPs was 1.42-fold higher than that in l-SPs when they interacted with α-syn aggregates (Figures 4a and 4b). The XPS results showed that when l/d-SPs interacted with α-syn under NIR irradiation, more Cu2+ and Co3+ were produced than were shown by the XPS data for the single l/d-SPs (Figures 1e and 1f and Supporting Information Figures S1e and S1f, S27, and S28). Based on these XPS results, the ratios of Cu+/Cu2+ and Co2+/Co3+ in l/d-SPs were analyzed in Supporting Information Tables S2 and S3. After reacting with α-syn under NIR irradiation, the ratio of Cu+/Cu2+ in d-SPs decreased from 0.26 to 0.21, and the ratio of Co2+/Co3+ ranged from 1.91 to 0.74; while the ratio of Cu+/Cu2+ in l-SPs decreased from 0.25 to 0.22, and the ratio of Co2+/Co3+ ranged from 1.53 to 1.34. Compared with Cu+/Cu2+ ratio, Co2+/Co3+ ratio had a great influence on surface oxygen defects. The ratio of Co2+/Co3+ in d-SPs was much lower than that of l-SPs under the same conditions, indicating that d-SPs disaggregated α-syn more efficiently. Figure 4 | ROS production of the (a) d-SPs and (b) l-SPs under different irradiation time (0, 3, 5, and 10 min). (c) Depolymerization kinetics of α-syn fibrillization after incubation with or without l/d-SPs in the same reaction time monitored by ThT fluorescence. (d) Depolymerization kinetics of α-syn fibrillization with or without ROS inhibitor using aliquots of the reaction monitored by ThT fluorescence. Data are presented as mean ± s.d. (n = 5). Download figure Download PowerPoint We also selected a ROS inhibitor (N-acetyl-l-cysteine, NAC) for the control experiment, and ROS production was not detected under the same conditions in the d-SPs and α-syn aggregate mixtures when NAC was added ( Supporting Information Figure S29). In addition, the thioflavin-T (ThT) binding assay demonstrated that the d-SPs had better ability to dissociate fibrils after 10 min of NIR irradiation (Figure 4c). The results of the ThT fluorescence assay showed that the α-syn fibrils disaggregated into short fragments and oligomers with increased illumination time. In the group with NAC added, there was almost no change in α-syn fibrils (Figure 4d). These results indicate that d-SPs induced the generation of ROS, which significantly accelerated the dissociation of α-syn fibrils. d-SPs prevented α-syn aggregation in the PD model mice In addition to good biocompatibility, the d-SPs had a high inhibitory effect on α-syn fibrillization ( Supporting Information Figure S30), which prompted us to study their therapeutic effect on PD in vivo. PD was induced in mice by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP).25 The d-SPs (20 μg) were injected into the brain of MPTP-treated PD mice. Mice (n = 5) were irradiated for 10 min daily for 60 days (Figure 5a). To evaluate the therapeutic effect of the d-SPs on the PD model mice, we optimized the NIR irradiation time and analyzed brain striatum tissue sections from the mice with which we used immunofluorescence staining (Figure 5b). Compared with the wild-type (WT) group, the PD group showed decreased expression of tyrosine hydroxylase (Th) and increased expression of the ionized calcium binding adaptor molecule-1 (Iba-1) and α-syn protein.23 After d-SPs treatment, the expression of Iba-1 and α-syn protein in the substantia nigra of mice was significantly reduced, while the expression of Th increased. The levels of α-syn in PD mice were almost the same as those in the WT group after 10 min exposure. The corresponding level of Iba-1 decreased with increased light duration per day (Figures 5c–5e). The concentration of α-syn in the cerebrospinal fluid was also measured using an enzyme-linked immunosorbent assay (ELISA) kit. After 2 months of treatment, compared with the α-syn concentration in the cerebrospinal fluid of the PD group, the α-syn concentration of mice in the 5-min light group decreased to 4.92 ng/mL, the 10-min light group decreased to 1.022 ng/mL, and the 20-min light group dropped to 0.91 ng/mL. Analysis of the data showed that the daily α-syn level in the 10-min light group was very close to the α-syn level in the WT group, which decreased about 67.5% compared with the PD group (Figure 5f and Supporting Information Figure S31). Figure 5 | (a) Schematic illustration of d-SPs injected into the PD mice. (b) Confocal images of wild mice striatum, PD mice striatum, and treated PD mice striatum, immunostained for Iba-1, Th, and α-syn. Red: first column of Iba-1 and second column of Th, Green: α-syn. Scale bar, 50 μm. Average expression levels of each treatment group (c) Iba-1, (d) Th, and (e) α-syn. (f) α-Syn concentration in cerebrospinal fluid of PD mice after treatment measured by ELISA. Data are presented as mean ± s.d. (n = 5). ***p < 0.001, analyzed by Student’s t-test. Download figure Download PowerPoint Immunohistochemistry (IHC) was performed to observe the distribution of Th and α-syn in the striatum of mice. As shown in Figure 6a, Supporting Information Figures S32 and S33, the Th content in the treatment group was 6.07-times higher than that in the PD group, and the α-syn content was 3.53-times lower than that in the PD group, which was consistent with the immunohistofluorescence (IHF) results shown in Figure 5b. The same results are also shown on the confocal microscope image (Figure 6b). Furthermore, histological sections of the heart, liver, spleen, lung, kidney, and substantia nigra stained with hematoxylin and eosin (H&E) showed no significant organ damage or inflammation, indicating that the mice in the treatment group had no obvious adverse reactions after long-term exposure to d-SPs ( Supporting Information Figures S34 and S35). Figure 6 | (a) IHC images of wild mice striatum, PD mice striatum, and treated PD mice striatum, immunostained for Th and α-syn. Scale bar, 100 μm. (b) IHF images of wild mice brains, PD mice brains, and treated PD mice brains, immunostained for Th and α-syn. Scale bar, 1 mm. (c) The latent period in water maze of the wild mice and PD mice to find the target quadrant after treatment. (d) The time in the target quadrant of the wild mice and PD mice after treatment. (e) The crosses times in target quadrant of the wild mice and PD mice after treatment. (f) The track sheets of the wild mice and PD mice after treatment (n = 5 mice per group). **p < 0.01; ***p < 0.001, analyzed by Student’s t-test. Download figure Download PowerPoint In addition, photoacoustic imaging was performed on mice injected with d-SPs stereotaxically in the brain to monitor the changes of d-SPs under the illumination of NIR within 60 days. After injection of d-SPs, it was found that the photoacoustic signal gradually weakened with the extension of time, and the signal became extremely weak on the 60th day, which was basically consistent with the control group, indicating that d-SPs could be excreted from the mice brain ( Supporting Information Figure S36), and it was further confirmed from the photoacoustic imaging of d-SPs in tissues (brain, liver, and kidney) taken from mice treated at 20, 40, and 60 days. On the 20th day, obvious signals could be seen in the liver and kidney tissues, and the signal strength gradually decreased with time. These data showed that d-SPs could be excreted from the brain and enriched in the liver and kidney. In addition, the changes in the concentration of Cu and Co in the blood of mice was analyzed by inductively coupled plasma mass spectrometry (ICP-MS) every 15 days for a total of 60 days after in situ injection of d-SPs. It was found that the concentration of Cu and Co gradually increased, reaching a peak on the 30th day. On the 60th day, the concentration of the two elements was similar to the control group. TEM and DLS spectra of d-SPs in cell culture mediums after illumination exhibited that the SPs gradually dissolved into small nanoparticles ( Supporting Information Figure S37). The Morris water maze (MWM) test was then used to evaluate the spatial cognition and memory of model mice. The mice were trained daily for 5 days to find the hidden platform. The platform was removed on the 6th day for probe testing. Compared with WT mice, PD model mice showed obvious defects, such as longer latency and shorter time spent in the target quadrant after removing the platform (Figures 6c and 6d). There were fewer passes through the stage and more random movement (Figures 6e and 6f). In the 5-day training and subsequent exploratory experiments, compared with PD model mice, the d-SPs-treated mice showed significant improvements in spatial learning and memory. Combined with the above results, d-SPs effectively reduced α-syn aggregate deposition in the brain and restored memory function in PD mice. Conclusion In this study, chiral CuxCoyS l/d-SPs modified with l- or d-Pen were prepared. These chiral nanoparticles, especially the d-SPs, effectively inhibited α-syn aggregation and had the ability to cause α-syn fibrils disaggregation under NIR irradiation. NIR illumination activated d-SPs, which generated a large number of ROS, destroyed the stability of the β-sheet conformation of α-syn fibrils, prevented the fibrillation of monomers, and reduced their cytotoxicity. In vivo studies of PD model mice showed that the chiral SPs successfully removed α-syn fibrils deposited in the brain, and the SPs decomposed into small nanoparticles and were excreted through the blood–brain barrier. Based on the effectiveness of this treatment in the PD model mice, we hypothesize that the NIR light-activated PD treatment strategy will provide opportunities for the application of chiral inorganic SPs in the treatment of protein aggregation diseases such as neurodegeneration diseases. Supporting Information Supporting Information is available and includes detailed experimental methods, supplementary figures and tables regarding SPs analysis, kinetics of α-syn fibrils disaggregation, cell viability measurement, confocal images, and ROS detection. Conflict of Interest There is no conflict of interest to report. Funding Information This work is financially supported by the National Natural Science Foundation of China (nos. 32071400, 21977038, 51902136, and 21874058) and the Fundamental Research Funds for the Central Universities (no. JUSRP12003). References 1. Schapira A. H. V.; Olanow C. W.; Greenamyre J. T.; Bezard E.Slowing of Neurodegeneration in Parkinson’s Disease and Huntington’s Disease: Future Therapeutic Perspectives.Lancet2014, 384, 545–555. Google Scholar 2. Abeliovich A.; Gitler A. D.Defects in Trafficking Bridge Parkinson’s Disease Pathology and Genetics.Nature2016, 539, 207–216. Google Scholar 3. Ma W.; Xu L.; de Moura A. F.; Wu X.; Kuang H.; Xu C.; Kotov N. A.; Chiral Inorganic Nanostructures.Chem. Rev.2017, 117, 8041–8093. Google Scholar 4. Li F.; Li Y.; Yang X.; Han X.; Jiao Y.; Wei T.; Yang D.; Xu H.; Nie G.Highly Fluorescent Chiral N-S-Doped Carbon Dots from Cysteine: Affecting Cellular Energy Metabolism.Angew. Chem. Int. Ed. Engl.2018, 57, 2377–2382. Google Scholar 5. Li Y.; Cheng J.; Li J.; Zhu X.; He T.; Chen R.; Tang Z.Tunable Chiroptical Properties from the Plasmonic Band to Metal-Ligand Charge Transfer Band of Cysteine-Capped Molybdenum Oxide Nanoparticles.Angew. Chem. Int. Ed. Engl.2018, 57, 10236–10240. Google Scholar 6. Jiang S.; Chekini M.; Qu Z. B.; Wang Y.; Yeltik A.; Liu Y.; Kotlyar A.; Zhang T.; Li B.; Demir H. V.; Kotov N. A.Chiral Ceramic Nanoparticles and Peptide Catalysis.J. Am. Chem. Soc.2017, 139, 13701–13712. Google Scholar 7. Zhang H.; Hao C.; Qu A.; Sun M.; Xu L.; Xu C.; Kuang H.Light-Induced Chiral Iron Copper Selenide Nanoparticles Prevent Beta-Amyloidopathy in Vivo.Angew. Chem. Int. Ed. Engl.2020, 59, 7131–7138. Google Scholar 8. Hao C.; Qu A.; Xu L.; Sun M.; Zhang H.; Xu C.; Kuang H.Chiral Molecule-Mediated Porous CuxO Nanoparticle Clusters with Antioxidation Activity for Ameliorating Parkinson’s Disease.J. Am. Chem. Soc.2019, 141, 1091–1099. Google Scholar 9. Hao C.; Wu X.; Sun M.; Zhang H.; Yuan A.; Xu L.; Xu C.; Kuang H.Chiral Core-Shell Upconversion [email protected] Nanoassemblies for Quantification and Bioimaging of Reactive Oxygen Species in Vivo.J. Am. Chem. Soc.2019, 141, 19373–19378. Google Scholar 10. Qu A.; Sun M.; Kim J. Y.; Xu L.; Hao C.; Ma W.; Wu X.; Liu X.; Kuang H.; Kotov N. A.; Xu C.Stimulation of Neural Stem Cell Differentiation by Circularly Polarized Light Transduced by Chiral Nanoassemblies.Nat. Biomed. Eng.2021, 5, 103–113. Google Scholar 11. Hou K.; Zhao J.; Wang H.; Li B.; Li K.; Shi X.; Wan K.; Ai J.; Lv J.; Wang D.; Huang Q.; Wang H.; Cao Q.; Liu S.; Tang Z.Chiral Gold Nanoparticles Enantioselectively Rescue Memory Deficits in a Mouse Model of Alzheimer’s Disease.Nat. Commun.2020, 11, 4790. Google Scholar 12. Park J. I.; Nguyen T. D.; de Queiros Silveira G.; Bahng J. H.; Srivastava S.; Zhao G.; Sun K.; Zhang P.; Glotzer S. C.; Kotov N. A.Terminal Supraparticle Assemblies from Similarly Charged Protein Molecules and Nanoparticles.Nat. Commun.2014, 5, 3593. Google Scholar 13. Yang M.; Chan H.; Zhao G.; Bahng J. H.; Zhang P.; Kral P.; Kotov N. A.Self-Assembly of Nanoparticles into Biomimetic Capsid-like Nanoshells.Nat. Chem.2017, 9, 287–294. Google Scholar 14. Maiolo D.; Pigliacelli C.; Sanchez Moreno P.; Violatto M. B.; Talamini L.; Tirotta I.; Piccirillo R.; Zucchetti M.; Morosi L.; Frapolli R.; Candiani G.; Bigini P.; Metrangolo P.; Baldelli Bombelli F.Bioreducible Hydrophobin-Stabilized Supraparticles for Selective Intracellular Release.ACS Nano2017, 11, 9413–9423. Google Scholar 15. Zhou Y.; Marson R. L.; van Anders G.; Zhu J.; Ma G.; Ercius P.; Sun K.; Yeom B.; Glotzer S. C.; Kotov N. A.Biomimetic Hierarchical Assembly of Helical Supraparticles from Chiral Nanoparticles.ACS Nano2016, 10, 3248–3256. Google Scholar 16. Yeom J.; Guimaraes P. P. G.; Ahn H. M.; Jung B. K.; Hu Q.; McHugh K.; Mitchell M. J.; Yun C. O.; Langer R.; Jaklenec A.Chiral Supraparticles for Controllable Nanomedicine.Adv. Mater.2019, 32, 1903878. Google Scholar 17. Kim J. Y.; Yeom J.; Zhao G.; Calcaterra H.; Munn J.; Zhang P.; Kotov N.Assembly of Gold Nanoparticles into Chiral Superstructures Driven by Circularly Polarized Light.J. Am. Chem. Soc.2019, 141, 11739–11744. Google Scholar 18. Yang Y.; Wang B.; Shen X.; Yao L.; Wang L.; Chen X.; Xie S.; Li T.; Hu J.; Yang D.; Dong A.Scalable Assembly of Crystalline Binary Nanocrystal Superparticles and Their Enhanced Magnetic and Electrochemical Properties.J. Am. Chem. Soc.2018, 140, 15038–15047. Google Scholar 19. Zhao Y.; Huang Y.; Zhu H.; Zhu Q.; Xia Y.Three-in-One: Sensing, Self-Assembly, and Cascade Catalysis of Cyclodextrin Modified Gold Nanoparticles.J. Am. Chem. Soc.2016, 138, 16645–16654. Google Scholar 20. Li S.; Liu J.; Ramesar N. S.; Heinz H.; Xu L.; Xu C.; Kotov N. A.Single- and Multi-Component Chiral Supraparticles as Modular Enantioselective Catalysts.Nat. Commun.2019, 10, 4826. Google Scholar 21. Feng W.; Kim J. Y.; Wang X.; Calcaterra H. A.; Qu Z.; Meshi L.; Kotov N. A.Assembly of Mesoscale Helices with Near-Unity Enantiomeric Excess and Light-Matter Interactions for Chiral Semiconductors.Sci. Adv.2017, 3, e1601159. Google Scholar 22. Gao G.; Chen R.; He M.; Li J.; Li J.; Wang L.; Sun T.Gold Nanoclusters for Parkinson’s Disease Treatment.Biomaterials2019, 194, 36–46. Google Scholar 23. Kim D.; Yoo J. M.; Hwang H.; Lee J.; Lee S. H.; Yun S. P.; Park M. J.; Lee M.; Choi S.; Kwon S. H.; Lee S.; Kwon S. H.; Kim S.; Park Y. J.; Kinoshita M.; Lee Y. H.; Shin S.; Paik S. R.; Lee S. J.; Lee S.; Hong B. H.; Ko H. S.Graphene Quantum Dots Prevent Alpha-Synucleinopathy in Parkinson’s Disease.Nat. Nanotechnol.2018, 13, 812–818. Google Scholar 24. Yoo J.; Lee E.; Kim H. Y.; Youn D. H.; Jung J.; Kim H.; Chang Y.; Lee W.; Shin J.; Baek S.; Jang W.; Jun W.; Kim S.; Hong J.; Park H. J.; Lengner C. J.; Moh S. H.; Kwon Y.; Kim J.Electromagnetized Gold Nanoparticles Mediate Direct Lineage Reprogramming into Induced Dopamine Neurons in Vivo for Parkinson’s Disease Therapy.Nat. Nanotechnol.2017, 12, 1006–1014. Google Scholar 25. Singh N.; Savanur M. A.; Srivastava S.; D’Silva P.; Mugesh G.A Redox Modulatory Mn3O4 Nanozyme with Multi-Enzyme Activity Provides Efficient Cytoprotection to Human Cells in a Parkinson’s Disease Model.Angew. Chem. Int. Ed. Engl.2017, 56, 14267–14271. Google Scholar Previous articleNext article FiguresReferencesRelatedDetails Issue AssignmentVolume 4Issue 7Page: 2440-2451Supporting Information Copyright & Permissions© 2021 Chinese Chemical SocietyKeywordschiralityCuxCoySα-synucleinParkinson’s diseasesupraparticles Downloaded 852 times PDF DownloadLoading ...